Vous êtes sur la page 1sur 17

Molecular Ecology (2001) 10, 1855 1871

REVIEW ARTICLE
Blackwell Science, Ltd

The molecular revolution in ectomycorrhizal ecology: peeking into the black-box


T H O M A S R . H O RT O N * and T H O M A S D . B R U N S *Department of Forest Science, Oregon State University, Corvallis, OR, 97331, USA, Department of Plant and Microbial Biology, 111 Koshland Hall, University of California, Berkeley, CA 94720, USA

Abstract
Molecular tools have now been applied for the past 5 years to dissect ectomycorrhizal (EM) community structure, and they have propelled a resurgence in interest in the eld. Results from these studies have revealed that: (i) EM communities are impressively diverse and are patchily distributed at a ne scale below ground; (ii) there is a poor correspondence between fungi that appear dominant as sporocarps vs. those that appear dominant on roots; (iii) members of Russulaceae, Thelephoraceae, and/or non-thelephoroid resupinates are among the most abundant EM taxa in ecosystems sampled to date; (iv) dissimilar plants are associated with many of the same EM species when their roots intermingle this occurs on a small enough spatial scale that fungal individuals are likely to be shared by dissimilar plants; and (v) mycoheterotrophic plants have highly specic fungal associations. Although, these results have been impressive, they have been tempered by sampling difculties and limited by the taxonomic resolution of restriction fragment length polymorphism methods. Minor modications of the sampling schemes, and more use of direct sequencing, has the potential to solve these problems. Use of additional methods, such as in situ hybridization to ribosomal RNA or hybridization coupled to microarrays, are necessary to open up the analysis of the mycelial component of community structure.
Keywords: community ecology, ectomycorrhizae, fungi, PCR Received 7 January 2001; revision received 19 April 2001; accepted 19 April 2001

Introduction
Ectomycorrhizal (EM) symbioses are important on a global scale because the dominant trees in most of the worlds temperate and boreal forests and in large areas of tropical and subtropical forests are ectomycorrhizal (Allen 1991; Read 1991). For host trees these fungi serve as the primary nutrient gathering interface, scavenging nitrogen, phosphorus, and trace elements from both inorganic and organic pools in the soil; this is done in exchange for xed carbon from the tree. For many trees, such as those in the Pinaceae, Fagaceae, Betulaceae, and Dipterocarpaceae, it is clear that normal growth and survival is dependent on colonization by EM fungi (Smith & Read 1997). EM fungi are also important because they include high value
Correspondence: Thomas R. Horton. Present address: 350 Illick Hall, SUNY-ESF, Syracuse, NY, 13210, USA. Fax: +1 315 470 6934; E-mail: trhorton@esf.edu 2001 Blackwell Science Ltd

edible fungi such as trufes (Tuber spp.), matsutake (Tricholoma spp.), chanterelles (Cantharellus spp.), and Boletus edulis. Fungi that form ectomycorrhizae are not a monophyletic group; traditional classication and recent molecular systematic studies demonstrate that this trait has been independently derived many times and perhaps convergently lost as well (Gargas et al. 1995; Hibbett et al. 1997, 2000 Bruns et al. 1998). Over 5000 species of EM fungi have been described (Molina et al. 1992). The symbionts span all of the phyla of true fungi (Zygomycota, Ascomycota, and Basidiomycota), and occur in at least 15 families within the Basidiomycota and Ascomycota. However, from molecular clock estimates and the fossil records of their plant hosts, it appears that most of the diversity arose relatively recently, perhaps initially in the Cretaceous with later extensive radiation around 2560 Ma (LePage et al. 1997; Bruns et al. 1998). This pattern is in contrast to that of the arbuscular mycorrhizal fungi that fall within a single

1856 T. R . H O RT O N and T. D . B R U N S monophyletic group, the Glomales, and appear to have arisen with the land plants 400 500 Ma (Simon et al. 1993; Pirozynski & Dalp 1989). EM fungus communities are species rich, and many basic questions about these communities remain unanswered. These include: (i) what is the structure of these communities in terms of numbers and abundance of species; (ii) how is structure maintained and affected by factors such as host diversity, soil types, organic inputs, disturbance, and succession; (iii) does structure have signicant effects on ecosystem functions (e.g. nutrient cycling and retention), or on plant community structure; and (iv) what is the natural history or autecology of the dominant species. Tackling these questions in a quantitative manner presents several challenges unique to mycorrhizal fungi. The rst problem is that EM species composition is not easily manipulated in natural or laboratory settings. This is because a large percentage of EM fungi either grow poorly or do not grow at all in culture, and because addition of spores or mycelial inoculum rarely results in colonization under nonsterile conditions. Seedlings that are inoculated under sterile or semi-sterile conditions can be out-planted, but replacement by indigenous fungi is a common result (e.g. Bledsoe et al. 1982; Danielson & Visser 1989), except perhaps when the host is planted outside its natural range (Selosse et al. 1998a). Much has been learned about behaviour and physiological ecology of individual species, particularly of Suillus, Rhizopogon, Paxillus, Laccaria, Pisolithus, and Cenococcum, in laboratory microcosms, but the overwhelming majority of EM fungi have not been amenable to such manipulations. Second, vegetative structures of these fungi (i.e. mycorrhizae and mycelium in the soil) occur largely below ground and are difcult to track and identify. Attaching species names to below ground structures has been problematic because the taxonomy of these organisms is based on their sexual states (e.g. mushrooms, trufes, etc.), and the vegetative states are much smaller and more difcult to distinguish. For these reasons most of the earlier literature on EM communities was based on collecting and quantifying fruiting structures, or, if based on mycorrhizae, left large numbers of species unidentied or lumped into nebulous types (e.g. brown type). Thus, the below ground dynamics and roles of the fungi were largely relegated to a single functional guild, if not the so-called black-box (Allen 1991). & Faloona 1987; Gardes et al. 1991; Henrion et al. 1992; Lanfranco et al. 1998). The primary amplication targets have been ribosomal genes and spacers; these regions combine the advantages of high copy number, highly conserved sequence tracks that can serve as sites for primer design, and variable regions between the priming sites. Both universal, and fungal or plant specic primers, have been designed (White et al. 1990; Cullings 1992; Gardes & Bruns 1993; Egger 1995; Vrlstad et al. 2000), the former have been designed for various taxonomic levels. The use of PCR and ribosomal genes is a familiar theme in microbial ecology. The eld of EM community ecology differs from other areas of microbial ecology in that these methods are used almost strictly for identication, rather than for both identication and quantication, as is common in prokaryotic communities. This is a tremendous advantage, because it avoids most of the sticky issues and additional effort involved with PCR quantication. The reason molecular quantication is generally unnecessary is that ectomycorrhizae are small macroscopic packages (e.g. colonized root tips) that can be counted or weighed. In contrast to the situation in arbuscular mycorrhizae, EM colonization is usually obvious from the external appearance of the root tips (Fig. 1). A few taxa of fungi can be more cryptic in their colonization, such as the so-called dark septate fungi (Phialophora, Chloridium and Pialocephela) or other ascomycetous fungi such as Wilcoxina spp., but even these can be recognized with a little effort. In addition to the advantage of independent quantication, morphological differences among ectomycorrhizae can be used to sort species. Some differences such as shape and colour are obvious even to the untrained eye and can be sorted into discrete groups rapidly but are insufcient to differentiate among closely related species (Fig. 1). Other morphological characters, such as those that involve hyphal types and arrangements, require signicant experience to recognize and more careful treatment of the specimens, but can yield identications directly without molecular analyses in systems that are well characterized (Agerer 198796; Ingleby et al. 1990; Goodman et al. 1996 98). There are trade-offs involved here, however. The more thorough the morphological characterization, the fewer the samples that can be sorted in a given time; therefore, the large numbers of samples necessary for communitylevel studies often makes a strictly morphological approach impractical for large scale questions. In addition, the DNA in samples degrades rapidly, so the longer the pre-extraction processing, the higher the risk that they will not amplify. Conversely, the cruder the sorting the more likely it is that a morphotype will contain multiple species, and thus defeat the purpose of the sorting. This raises the question of how much to morphotype, but we defer our discussion of this question until other sampling issues are addressed.
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

A marriage of molecules and morphology to the rescue


The most important methodological advance in the study of EM communities has been the application of the polymerase chain reaction (PCR) for identication (Mullis

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1857
Fig. 1 Variation in EM root tip morphology. (a) Hebeloma crustiliniforme and Douglas-r by R. Molina. (b) Cortinarius sp. and Douglasr by B. Zak. (c) Melanogaster intermedius and Arbutus menziesii by R. Molina. (d) Amanita muscaria and Picea sitchensis by R. Molina. (e) Rhizopogon vinicolor and Douglas-r, cross section of a tubercle by D. Luoma. (f) Lactarius rubrilacteus and Douglas-r by B. Zak. Scale bars equal 5 mm.

RFLP matching analysis is king, but the monarchy is insufcient Fungus identication with ITS
Most of the molecular ecology on EM fungi has involved restriction analyses of the internal transcribed spacer (ITS) region. This nuclear region, which is well known to the elds of molecular ecology and fungal systematics, lies between the small subunit (SSU) and the large subunit (LSU) ribosomal RNA (rRNA) genes and contains two noncoding spacer regions separated by the 5.8S rRNA gene. In fungi it is typically about 650900 bp in size, including the 5.8S gene. It is usually amplied by either the universal primer pair (ITS1 and ITS4; White et al. 1990; Gardes et al. 1991), or a fungal specic, or basidiomycete specic pair (ITS1f and ITS4 or ITS1f and ITS4b, respectively; Gardes & Bruns 1993). There has been a mistaken impression in the literature that the ITS1 and ITS4 primers are fungal specic rather than universal; this idea has been reinforced by the observation that they do not amplify the ITS of the Pinaceae very well. Nevertheless, ITS1 and ITS4 were designed with plant sequences in mind and in some hosts (e.g. members of the Monotropoideae) these sequences are co-amplied with those of the target
2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

fungi. Many other taxon-specic ITS primers have been developed. However, primers designed to show specicity for a particular group should be used with caution as a lack of amplication with the primer pair can be misinterpreted. For instance, primers intended for a specic group (e.g. Basidiomycete or Ascomycete) may not amplify DNA from every species within the intended group. Restriction fragment length polymorphism (RFLP) analysis of the ITS region has been popular because it separates many species quickly and relatively cheaply with minimal technical requirements. Typically two or three restriction enzyme digests are enough to distinguish most species (Nylund et al. 1995; Gardes & Bruns 1996a; Dahlberg et al. 1997; Krn et al. 1997; Pritsch et al. 1997, 2000; Gehring et al. 1998; Jonsson et al. 1999a,b; Mahmood et al. 1999; Eberhardt et al. 2000; Methvyn et al. 2000). The reason so few digests are needed is that sequence differences between taxa are usually the result of indels, insertion or deletions of nucleotides that cause length variations. Thus, it is not necessary for enzyme recognition sites to change in order to create a unique RFLP pattern; enzymes that cleave the region into small fragments will reveal small size differences. For these reasons side-by-side comparisons of patterns is an extremely simple and sensitive method for determining identity or near identity at the species level,

1858 T. R . H O RT O N and T. D . B R U N S and should be employed to check putative matches observed from different gels. The main problem with the ITS RFLP approach is that when it is used alone, the number of unidentied types typically remains quite high. This is true even at locations from which extensive ITS RFLP databases are available from sporocarp samples (Krn & Nylund 1997). This is partially caused by the fact that RFLP databases tend to be composed primarily of species that make large or obvious sporocarps, and as discussed below, these species are often not the most common ones encountered on colonized roots. However, even when a species is represented in an RFLP database, matching an unknown to it presents additional problems. Exact matches based on database searches are uncommon, because size estimates for fragments vary and intraspecic variation exists across large geographical scales (Krn et al. 1997; Farmer & Sylvia 1998; Selosse et al. 1998a; Methvyn et al. 2000). Furthermore, the meaning of near matches cannot be satisfactorily quantied, because the information content of an RFLP pattern is very limited. For example, when ITS RFLP patterns are used to build phylogenetic trees, different species within a genus are frequently dispersed across the tree, and branch support metrics, such as bootstrap values, are either not employed or are below levels that one would desire for condent placements. Switching to acrylamide gels, automated sequencers, and genescan technology solves many of the accuracy and comparison problems (Peter et al. 2001), but increases the cost/sample substantially. An additional problem with RFLP databases is that they are not standardized in terms of the primers used, the enzymes used, or the way the information is stored and retrieved. They are also problematic in that there is no standard location to nd such data, although some web sites are under construction. We suggest that RFLP databases work best when they are focused on specic ecosystems, which will keep the fungal diversity low enough to minimize intraspecic variation in ITS sequences, thereby increasing the chances of matches and reducing the potential for over estimating diversity. A regional focus will also keep the fungal diversity low enough to allow one to master the morphological variation of EM. However, the central problem of unknown fungal RFLP patterns is not easily overcome by RFLP approaches alone and direct sequence analysis is required for the identication of many unidentied RFLP types (see below). subjected to RFLP analysis, and compared to patterns from unknown samples to those generated from identied leaf extracts (Horton & Bruns 1998; Horton et al. 1999; Stendell et al. 1999; Cullings et al. 2000). We have had good success separating conifers at the genus level (Pinus, Pseudotsuga, Tsuga). Angiosperms pose a more difcult situation in that, in our experience, some are difcult to amplify with this primer pair (Quercus/Fagaceae, Adenostoma/Rosaceae, Pickeringia/Fabaceae). Developing plant specic primers for the ITS region or another target that allows specic amplication of a wider variety of plant species and allows separation of species is desirable.

The addition of sequence analysis to get beyond unknowns


An important step to take for at least the dominant RFLP types identied in a given study is direct sequence analysis, but relatively few researchers have done so. The additional effort or expense required to generate sequence data has lessened to the point that obtaining sequence data is as routine as obtaining RFLP data. While there are an increasing number of loci or genetic regions available for sequence analysis, currently several options are particularly useful. For most microbial ecologists the molecule of choice would typically be the SSU rRNA gene. For EM fungi this gene has not been a target, partially because many of the critical taxa have not been sequenced and more importantly because the resolution is low relative to the amount of sequencing effort required (Bruns et al. 1992). However, 18S sequences have now been determined for some of the important ectomycorrhizal groups such as the Pezizales, Gomphaceae, and Cantharellaceae (Hibbett et al. 1997; Landvik et al. 1997; ODonnell et al. 1997; Norman & Egger 1999; Pine et al. 1999), which should increase the value of this gene for identication of EM fungi. A very quick way to search for SSU sequence matches is via the Ribosomal Database project, which provides web-based search and treeing algorithms (http://www.cme.msu.edu/RDP/html/ index.html; Olsen et al. 1993). The 5 end of LSU rRNA gene is another possible target for sequence identication. It is a more variable, and therefore a more informative, target than the SSU, but so far it appears to have been used only twice for identication of unknown EM ascomycetes. Here it allowed placement at the generic level or above (Baar et al. 1999; Taylor & Bruns 1999). It has now become much more useful as the large number of new sequences from the Agaricales and Boletales have recently been added (Moncalvo et al. 2000). A small piece of the mitochondrial large subunit rRNA gene (mtLSU) has been used extensively by our laboratory groups for identication of EM basidiomycetes (Bruns et al. 1998; Table 1). This region provides unambiguous
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

Plant identication with nuclear LSU


Some researchers may wish to identify the plant species from mycorrhizal root extracts. 28KJ is a plant specic primer that, in combination with a universal primer 28C, will amplify a small portion of the 28S rRNA gene of plants even when fungal DNA is present (e.g. an extract obtained from an EM root) (Cullings 1992). The PCR product is then

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1859
Table 1 Proportion of ectomycorrhizal root tip biomass identied Species match Horton & Bruns 1998 Horton et al. 1999 Stendell et al. 1999 Bidartondo et al. 2000* Taylor & Bruns 1999 Hemlock/Douglas-r Mean 1 S.D. of biomass 0.762 0.225 0.252 0.164 0.840 0.360 0.442 0.27 Family ID 0.174 0.600 0.555 0.659 0.154 0.383 0.415 0.20 Unknown RFLP 0.064 0.122 0.186 0.116 0.015 0.004 0.084 0.06 Unknown morphology 0 0.052 0.007 0.061 0 0.217 0.055 0.08

*Without ECM directly associated with Sarcodes root ball (Bidartondo, personal communication). Proportion of ECM in samples from soils samples taken in the forest. Horton, unpublished data.

placements of unknown sequences into family sized groups such as Russulaceae, Thelephoraceae, Amanitaceae, Suillus and allies, Boletus and allies, Gomphus and allies, Hygrophoraceae, and Cantharellaceae. However, it provides little or no resolution within these groups. Initially, placements within Cortinariaceae and Tricholomataceae, two important EM fungal families, were difcult to interpret, but we have recently added a number of taxa from Tricholoma, Inocybe and Cortinarius and are condent that unknown sequences from these genera are now placed within separate clades. One important problem with this database is that the mitochondrial genome in fungi can contain introns, which make it difcult to amplify the region. The introns can be present in some, but not all collections of a species and although usually rare, they are common in a few genera (Bruns et al. 1998). Several primers have been developed that allow amplication around the introns, but these primers have not worked across all genera (e.g. Albatrellus). The 5.8S nuclear rRNA gene, which is included within the ITS region, has also been used for very broad level identication (Cullings & Vogler 1998). This database was designed to help researchers identify the source of their unknown ITS PCR amplications as fungal, plant, or animal DNA. Such a tool is useful for avoiding problems of analysing nontarget DNA from extracts (Camacho et al. 1997; Redecker et al. 1999). It can also be applied to resolve the phylum of fungi. However, many fungal sequences are not clearly placed into phylla, and those that are do not receive a high level of bootstrap support. Thus, interpretation of such analyses needs to be cautious. The spacers within the ITS region are probably the ideal sequences to use for identication because they have the resolving power to place unknowns to the species level or at least within a species group. A fast way to take advantage of the ITS data currently deposited is to search GenBank or EMBL using only the spacer sequences from unknown samples; sequence variation within the spacers is so high that only very closely related taxa are retrieved. This approach was recently used to identify the EM
2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

associates of Pisonia grandis and Corallorhiza as members of the Thelephora/Tomentella clade (Chambers et al. 1998; McKendrick et al. 2000). Because there are relatively few EM species that are the focus of detailed research, this approach rarely allows identication at the species level, only genus or species group, and there are many gaps in the taxonomic coverage. Therefore, unknown ITS sequences can still be relatively uninformative if they belong to under-represented groups. Nevertheless, if all researchers deposited ITS sequences from at least the major species found in their studies, this would greatly increase chances that these species would eventually be identied, and it would also increase the comparability of species lists across studies. Species level identications with ITS sequences require greater prior taxon sampling and more thorough phylogenetic analysis. Thus, this approach is limited to genera in which an ITS-based molecular systematic study has been conducted. Currently this is true for only a handful of ectomycorrhizal genera (e.g. Wilcoxina, Tricharina, Dermocybe, Suillus, Peziza, Hebeloma, Plicaria) (Egger 1995; Liu et al. 1995, 1997; Kretzer et al. 1996; Norman & Egger 1996, 1999; Aanen et al. 2000), but ITS sequence studies for other genera continue to be published. In order for these phylogenetic studies to be useful for identications, the sequence alignments must be readily available, so that unknown sequences can be easily added without having to realign all previous sequences. Many alignments have been made available on individual web sites, but deposition on Treebase (http://www.herbaria.harvard.edu/treebase/), which provides a central depository for all phylogenetic studies, is becoming more common and makes it easier to nd and use such data.

The current picture derived from such studies Diversity and distribution
EM fungal communities are impressively diverse, even in stands dominated by a single plant species (Danielson

1860 T. R . H O RT O N and T. D . B R U N S
4 Qian et al. 1998 Horton and Bruns 1998 3 Log (Abundance) Horton et al. 1999 Stendell et al. 1999 Mahmood et al. 1999 Kranabetter et al. 1998 Hagerman et al. 1999
Fig. 2 Log-normal distribution of species abundance from data obtained from published sources. In all cases there is an inverse relationship between rarity and abundance for EM fungi below ground.

10

20

30

40

50

60

Rank ECM Abundance


Table 2 Comparison of investigations of below ground EM community structure Age (years) 1241 30 35 35 > 50 100 100 200400 not given 3438 1 90 7090 350400 Area represented 8 stands at 16 ha 2025 m2 0.1 ha 125 m2 625 m2 500 m2 2500 m2 10 000 m2 3000 m2 625 m2 625 m2 21 000 m2 not given 6400 m2 Area sampled for sporocarps 200 m2/stand not given 19 m2 not surveyed not given 838 m2 not surveyed not surveyed not given not surveyed not surveyed 21 000 m2 1300 m2 not surveyed No. of soil samples (diameter) 15 (2.8 cm)/stand 120 (5 cm) 31 (10 cm) 25 (10 cm) 12 (10 cm) 10 (1.5 1.5 cm) 24 (5 cm) 10 (2.8 cm) 50 15 (10 cm) 29 seedlings 198 (5.5 cm) 80 (2.8 cm) 40 (4.6 cm) Actual surface area cored 92 cm2/stand 2352 cm2 2418 cm2 1963 cm2 780 cm2 23 cm2 942 cm2 61.5 cm2 not applicable 1178 cm2 not applicable 4704 cm2 492 cm2 664 cm2 No. of species from ECM data 18 24/stand (135 total) 21 > 20 16 > 40 25 > 50 24 51 20 7 200* 51 80

Stand type Scots pine (birch)1 Norway spruce2 Bishop pine3 Bishop pine/Douglas r4 Arctostaphylos/Douglas r5 Norway spruce6 Ponderosa pine7 Scots pine8 Pinyon pine9 Bishop pine10 Bishop pine11 Douglas r/western hemlock12 Norway spruce/Scots pine13 Red r14
1Jonsson

et al. 1999b; 2Krn & Nylund 1997; 3Gardes & Bruns 1996a; 4Horton & Bruns 1998; 5Horton et al. 1999; 6Dahlberg et al. 1997; et al. 1999; 8Jonsson et al. 1999a; 9Gehring et al. 1998; 10Taylor & Bruns 1999; 11Baar et al. 1999; 12Luoma et al. 1997; 13Jonsson et al. 2000; 14Bidartondo et al. 2000. *morphotypes.
7Stendell

1984; Visser 1995; Dahlberg et al. 1997; Pritsch et al. 1997; Gehring et al. 1998; Goodman & Trofymow 1998a,b; Kranabetter & Wylie 1998; Hagerman et al. 1999; Horton et al. 1999; Jonsson et al. 1999a,b; Stendell et al. 1999; Byrd et al. 2000). The distribution pattern of the fungi as sampled below ground leads to an inverse relationship between abundance and rarity (Fig. 2; Danielson 1984; Taylor & Alexander 1989; Visser 1995; Gardes & Bruns 1996a; Dahlberg et al. 1997; Krn & Nylund 1997; Jonsson et al. 1999a). This distribution pattern impacts the view of species richness because of the large number of rare types.

Most studies have presented data on species richness based on 30 or fewer soil samples, often covering less than 1 ha. With these sampling efforts, 50 or fewer species of fungi were observed below ground (Table 2). In forests dominated by Douglas r in southern Oregon (western North America), over 200 morphologically distinct EM were recorded in 198 soil samples taken over an area of about 2.1 ha (Luoma et al. 1997). The occurrence of the fungi at such a ne scale of patchiness below ground makes sampling a challenge. Indeed, the sampling of EM roots is typically inadequate to get a true picture of species
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1861
50

40 Number of Species Horton et al. 1999 Stendell et al. 1999 Douglas-fir and Hemlock Horton and Bruns 1998

30

20

10

0 0 10 20 Number of Soil Samples


Fig. 3 Species-area curves from four studies, with the number of soil samples taken used as a substitute for the area sampled (McCune & Mefford 1999). The curve reaches the asymptote only in the Horton & Bruns (1998) study. Second-order jackknife estimates from these data support the conclusion that the observed number of taxa (o) was similar to the estimated number of taxa (e) only in Horton & Bruns (1998): Horton & Bruns (1998) o = 16, e = 16.2; Stendell et al. (1999) o = 48, e = 71.6; Douglas-r and Hemlock o = 42, e = 87.5; Horton et al. (1999) o = 40, e = 69.2 (jackknife estimate = S + r1(2n 3)/n r2(n 2)squared/n(n 1)) where s = the observed number of species, r1 = the number of species occurring in exactly one sampling unit, r2 = the number of species occurring in exactly two sample units, and n = the number of sample units (Burnham & Overton 1979; Palmer 1991). This suggests that greater sampling effort was required in the other studies to better represent the diversity of EM fungi. Similar levels of diversity confound sampling efforts of plant species in tropical rain forests of Borneo (Mueller-Dombois & Ellenberg 1974).

30

40

richness (Fig. 3). If the number of soil samples has such a dramatic effect on observations of diversity, then identifying changes in richness in communities with high species richness remains somewhat ambiguous ( Jonsson et al. 1999a,b, c; Stendell et al. 1999). The distribution of mycorrhizae of many species is clustered, and most species typically occur in less than 10% of the soil samples taken (Gardes & Bruns 1996a; Flynn et al. 1998; (25%); Goodman & Trofymow 1998a,b; Horton et al. 1999; Taylor & Bruns 1999; Stendell et al. 1999; Bidartondo et al. 2000). Individual soil cores generally contain multiple species and adjacent root tips are frequently colonized by different species, although one or two species usually dominate in a core (Gardes & Bruns 1996a; Dahlberg et al. 1997; Pritsch et al. 1997; Goodman & Trofymow 1998a; Horton & Bruns 1998; Horton et al. 1999; Jonsson et al. 1999a,b; Stendell et al. 1999; Taylor & Bruns 1999). In several studies, the most abundant types by biomass occurred in only one or two soil cores, suggesting that shifting the sample location by a few centimetres can cause a dramatic shift in perception of species presence and abundance (Stendell et al. 1999; Horton et al. 1999; Bidartondo et al. 2000). For instance, notable differences were observed in both species assemblage and dominance when replicate samples were taken in consecutive
2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

years, with year one samples taken only 25 cm from year two samples in control plots (Stendell et al. 1999). The degree of patchiness in the below ground occurrence of EM fungi as highlighted by these studies can be predicted, in part, by the fact that fungi initially colonize isolated points along a root system, and proliferate locally through vegetative reproduction. Some fungi appear to be distributed at larger scales of patchiness than that indicated above. Stendell et al. (1999) report that a species of Russulaceae was abundant in two consecutive years of sampling, and although it occurred in six of 24 soil cores taken over the two years, it was limited to only one of their 1 m2 plots. Dahlberg et al. (1997), reported that Piloderma croceum was found in only half of their study plots and then only in ve neighbouring cores. Yet, Piloderma was so abundant in these samples that it ranked as the most abundant species overall, accounting for almost 20% of the root tips sampled. In pinyon pine stands characterized by isolated trees analogous to islands, single EM fungi dominated single trees, and the dominant fungi varied between trees (Gehring et al. 1998). This pattern may be inuenced by the dry climate limiting airborne spore dispersal in many years, or by the stand structure of Pinyon pine, which typically has scattered trees rather than a closed forest.

1862 T. R . H O RT O N and T. D . B R U N S Some species are more uniformly distributed. In California, Tomentella sublilacina has been found to be present and often numerically dominant in both coastal and montane pine forests (Gardes & Bruns 1996a; Horton & Bruns 1998; Stendell et al. 1999; Taylor & Bruns 1999). Tylospora brillosa exhibits a similar pattern in Northern European Picea forests (Erland 1995; Dahlberg et al. 1997; Flynn et al. 1998; Jonsson et al. 1999c). Cenococcum geophilum was the second most abundant type reported by Dahlberg et al. (1997), and in contrast to Piloderma, it was found in every soil core. In fact, there are virtually no studies where Cenococcum has not be found; it has one of the most distinctive morphotypes and its distribution pattern had already been observed in pre-molecular studies. Dahlberg et al. 1997; Luoma et al. 1997; Jonsson et al. 1999b, 2000). In one study, approximately 0.1% of the ground sampled for sporocarps was sampled for root tips and in four of the studies the gure was less than 0.05%. This does not include the temporal aspect of sporocarp surveys being spread out over a number of years with multiple trips taken, while the root-tip collections were conducted with single sampling dates at each location. It may be that if one could sample the above and below ground occurrence of EM fungi with equal intensity, a better correspondence between species lists would be found for many fungi. However, for some fungi, abundance differences between above and below ground views are not so easily explained. When a species fruits frequently and abundantly above ground, yet appears rare below ground, or conversely is found as a frequent and abundant root colonizer but exhibits little fruiting, sampling differences are an unlikely explanation. The best documented example of such a pattern is provided by Dahlberg et al. (1997), who report that the species which accounted for 70% of the annual fruiting biomass correspond to less than 30% of the colonized root tips. This pattern also extended to individual species of Cortinarius, which were dominant and fairly uniformly distributed fruiters, but were relatively minor components of the below ground community. Differential investment strategies, which are well known in plants (Chapin & Shaver 1985; Tilman 1994), could explain much of the observed pattern in EM fungi. For example, species that are abundant on roots but fruit rarely or in low abundance may simply invest more in vegetative growth and competition than in reproduction (Gardes & Bruns 1996a). This hypothesis would predict that those species, which fruit abundantly and are rare root colonizers, are weak vegetative competitors. However, at least one such species, Suillus pungens, can form large genets, which extend over hundreds of square meters; this is not what one would expect from a species that invests little in vegetative growth and competition (Bonello et al. 1998). The contradiction suggests that S. pungens may have a larger carbon budget than expected for the low number of roots it appears to colonize. Several hypotheses about how it may obtain such additional carbon have been put forth, but remain to be tested (Gardes & Bruns 1996a; Bonello et al. 1998).

Lack of correspondence between sporocarp and root-tip views of fungal dominance


Perhaps the most striking pattern in EM communities revealed by molecular techniques is that there is a poor correspondence between species that fruit abundantly and those that are abundant on roots (Mehmann 1995; Gardes & Bruns 1996a; Dahlberg et al. 1997; Krn & Nylund 1997; Gehring et al. 1998; Jonsson et al. 1999a,b). A few species show a fairly consistent correspondence between above and below ground occurrence including Amanita francheti (Gardes & Bruns 1996a), several hypogeous fungi (Luoma et al. 1997), and species of Lactarius (Luoma et al. 1997; T. Horton, personal observation). However, as a rule, most species that fruit abundantly at a study site are not observed as abundant EM and, most species found below ground are not well represented in the sporocarp record from a site. Could the discrepancy be exacerbated by incongruous sampling efforts above and below ground? Certainly the lack of correspondence between above and below ground species lists is at least partially caused by sampling differences. Sporocarp production for most species of fungi can be sporadic at best, and a fungus may not fruit abundantly, or at all, at a particular site during a study (Luoma 1991). It is also not surprising that many species that fruit at a site are not found in the relatively limited below ground record. As discussed above, the overwhelming majority of species are not abundant in the samples. In addition, even the most abundant species are typically found in less than 10% of the soil samples. In other words, even when sporocarps are collected, collecting the corresponding EM roots is not a trivial matter. It is important to realise that the effort required to measure a study plot for diversity by sporocarp production is very different than that by EM root tips, especially if one is attempting to use the two measures to document the fungal community equally. Table 2 includes ve studies in which a direct comparison was made between above and below ground views of EM fungi (Gardes & Bruns 1996a;

Dominance by Russulaceae, Thelephoraceae and non-thelephoroid resupinates


A number of studies show that members of Russulaceae, Thelephoraceae, and non-thelephoroid resupinates are among the most abundant and frequent taxa on EM roots in conifer communities in both Europe and North America (Erland 1995; Gardes & Bruns 1996a; Dahlberg et al. 1997; Kernaghan et al. 1997; Luoma et al. 1997; Sylvia & Jarstfer 1997; Flynn et al. 1998; Goodman & Trofymow 1998a;
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1863 Horton & Bruns 1998; Qian et al. 1998; Jonsson et al. 1999c; Stendell et al. 1999; Taylor & Bruns 1999; Bidartondo et al. 2000; Peter et al. 2001; Lilleskov et al. 2002). This was also true in a community dominated by the angiosperm Arctostaphylos glandulosa (Horton et al. 1999). The fact that species in the Russulaceae are dominant is not too surprising since Russula and Lactarius sporocarps are relatively common in most EM communities. The dominance of Tylospora was reported prior to molecular studies (Taylor & Alexander 1989), but this pattern is now reinforced by a large number of additional studies from previously unsampled ecosystems (Erland 1995; Dahlberg et al. 1997; Flynn et al. 1998; Jonsson et al. 1999c; Lilleskov et al. 2002). The identied taxa in the Thelephoraceae have been members of the genus Tomentella, which, like Tylospora, form resupinate sporocarps. These are thin crustlike structures, that are often found on woody debris or litter, and because of this habit, Tomentella spp. were at one time thought to be saprotrophic (Larsen 1968). The fruiting structures are cryptic and tend to be ignored in surveys of sporocarps; thus, contributing to the mismatch between above and below ground species lists. Why species in Russulaceae and resupinate taxa are so abundant in many communities is unknown, but it suggests that these fungi are particularly good competitors and play critical functional roles in EM communities. These two important EM groups both require focused taxonomic treatments. were largely colonized by fungi present as resistant propagules in the soil (Baar et al. 1999). This resistant propagule bank had a similar composition before (Taylor & Bruns 1999) and after the re (Baar et al. 1999), and was composed primarily of Rhizopogon, Wilcoxina, and Tuber species that were rare EM types in the prere community (Taylor & Bruns 1999). This behaviour has striking parallels to seed banks in plant communities. Interestingly, in both re adapted pine communities, suilloid fungi (Suillus and Rhizopogon) became dominant EM types during the early stages of community development. Studies have also been conducted to investigate the effects of ground res that allow trees and some EM roots to survive. Following res that did not completely burn off the litter layer, EM fungal communities showed little change in species richness but a reduction in species evenness in burned stands ( Jonsson et al. 1999b). In a study where the litter layer was completely burned off, the most abundant species were reduced to undetectable levels, thereby increasing species evenness in the stand (Stendell et al. 1999). In this latter study, it was suggested that EM types that were in high abundance before the re preferentially colonized the organic layer.

Fungal networks and plant succession


Mycorrhizal fungi directly and indirectly inuence plant community dynamics and succession (Perry et al. 1989; Allen 1991; Smith & Read 1997; van der Heijden et al. 1998; Horton et al. 1999). One way they inuence plant communities is through fungal carbon transfer between plant hosts. The use of radioactive labels and more recently, stable isotopes, have demonstrated that interhost connections occur in nature and that net transfer of carbon can occur between plant species (Bjrkman 1960; Simard et al. 1997a,b; McKendrick et al. 2000). The quintessential example of such transfers involves mycoheterotrophic plants; these are nonphotosynthetic plants that obtain their carbon from fungal associates. Those that are associated with ectomycorrhizal fungi, ultimately obtain their carbon from surrounding photosynthetic hosts that are connected to the same fungi. This has been known since Bjrkmans eldwork in 1960, and has been recently demonstrated under laboratory conditions (McKendrick et al. 2000). What was unappreciated prior to molecular ecology, is that these plants have highly specic fungal associations, such that each plant species only associates with a narrow range of closely related fungi. This pattern has now been demonstrated for two distantly related genera of orchids (Taylor & Bruns 1997), and several genera in the Monotropoideae (Cullings et al. 1996). In fact, among the mycoheterotrophs examined to date, there are no exceptions to the rule that they have specic associations (Kretzer et al. 2000a).

Response to disturbance
Airborne anthropogenic sources of nitrogen are now quantitatively important in many ecosystems (Vitousek et al. 1997) and several studies have highlighted the response of fungi to N deposition. High N deposition appears to have a dramatic effect on fruiting (Arnolds 1988; Brandrud 1995; Jonsson et al. 2000; Peter et al. 2001). Krn & Nylund (1997) found that in contrast to previous studies based on sporocarp records, data based on EM roots showed that nitrogen deposition did not appear to affect the number of species observed in a stand. However, nitrogen deposition was found to alter the species composition and reduce ne-root biomass (Krn & Nylund 1997; Peter et al. 2001). In contrast, Lilleskov et al. (2002) report that increasing nitrogen deposition reduced the species richness as measured by fungal occurrence on EM roots, with certain species disappearing and others becoming more abundant. Other studies have investigated the response of EM fungi to re. Visser (1995) reported that 75% of jack pine root tips were colonized by Suillus brevipes 6 years after a stand replacing re, and that there was an increase in species richness as stands aged to 41 years, with a concomitant decrease in dominance by Suillus. Bishop pine seedlings that established in the rst year after a stand replacing re
2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

1864 T. R . H O RT O N and T. D . B R U N S Photosynthetic plants typically associate simultaneously with a large number of unrelated ectomycorrhizal fungi, providing a great potential for connections with dissimilar plants through shared mycorrhizal networks. Based on data from sporocarp occurrence, pure culture synthesis experiments, and soil bioassays, it had been hypothesized that EM fungi may affect competitive interactions among ectomycorrhizal hosts (Kropp & Trappe 1982; Molina & Trappe 1982; Perry et al. 1989; Smith et al. 1995; Simard et al. 1997a; Massicotte et al. 1999). Until recently, eld data were lacking which allowed denitive statements about individual fungi colonizing multiple hosts. We now know from three settings that roots of dissimilar plants that overlap in space tend to be colonized by the same fungi (Horton & Bruns 1998; Horton et al. 1999; Cullings et al. 2000). In one case, fungi that formed mycorrhizae on both Pseudotsuga menziesii and Pinus muricata were in the majority below ground, but whether this impacted competitive interactions between these species was unclear (Horton & Bruns 1998). This pattern was observed again with Pinus contorta and Picea engelmannii (Cullings et al. 2000). A similar pattern was observed in a third case where preferential establishment of Pseudotsuga in Arctostaphylos patches had been observed, implying the EM interaction inuences plant succession (Horton et al. 1999). Whether carbon transfer occurred between Arctostaphylos and the shaded Pseudotsuga seedlings was not tested, but even without carbon transfer, it is a clear benet to seedlings to simply tap into a mycelial network already developed and supported by other plants (Newman 1988). However, net carbon transfer between Betula papyrifera and shaded Pseudotsuga via shared EM fungi has been demonstrated (Simard et al. 1997b). Taken together these examples show that shared networks are common and the potential for them to effect plant competitive interactions should not be ignored. Suillus grevillei, and Cortinarius rotundisporus individuals in a variety of settings (De La Bastide et al. 1994; Gyrta et al. 1997; Anderson et al. 1998; Bonello et al. 1998; Selosse et al. 1998a,b, 1999; Sawyer et al. 1999; Zhou et al. 1999; Redecker et al. 2001). Taken together these studies show that different species vary with respect to their temporal persistence and vegetative spread; some species are continuously recruited through sexual reproduction and spore dispersal even in undisturbed forest settings, while others appear to establish rarely or at an early stage in community development, and then spread vegetatively. Nevertheless, even the smallest individuals occupy soil volumes that include intermingling roots of multiple plant hosts. All of the population studies identied above utilized sporocarps as the source of DNA (in some cases cultures were made from sporocarp tissue). It is desirable to generate similar data utilizing mixed DNA extracts such as those from EM root tips or soil hyphae. As the molecular tools are rened, this comparison will be possible, as demonstrated by Kretzer et al. (2000b) who developed microsatellite primers for Rhizopogon vinicolor, allowing them to amplify fungal DNA from EM extracts. In addition, studies are needed that not only delineate clone boundaries, but also assess the effective population size and the extent of gene ow between populations of EM fungi, as has been done for some plant pathogens. Results from such studies will have direct applications in evolutionary and conservation biology of fungi.

Unresolved sampling issues Frequency vs. biomass


Dominance in the EM community has often been gauged by either frequency or abundance. The former simply being the proportion of samples that contain a given species (absolute constancy), and the latter often being either the proportion of total EM or proportion of total dry weight for a given species. In the discussion above we have treated the two as equivalent measures, but in fact they provide different views of the community structure. Frequency measures are biased towards those types that are common, even if they contribute little to the overall number or biomass of EM, while abundance measures are biased towards types that form massive clusters of root tips even though the fungus may be infrequently encountered. Just how different the frequency and abundance views are can be seen in Fig. 4. Several species, including Cenococcum geophilum (8), Rhizopogon parksii (5), and R. subcaerulescens (21), look much more important by frequency than by abundance; these species were common and dispersed, but colonized relatively few tips. Species such as Amanita gemmata (6), Cortinarioid 1 (7), and Boletoid 1 (17) were more clustered and contributed high biomass from a
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

Population level advances delineation of individuals


Our understanding of the autecology of individual EM species have been advanced in several ways by recent population level studies that have been directed toward identifying the physical extent of individual genotypes. Prior to the development of molecular tools for population studies, somatic incompatibility was used for this purpose on some of the few taxa that could be cultured. However, it was shown by Jacobson et al. (1993) that somatic incompatibility tests do not discriminate individuals as well as molecular approaches. Based on sporocarp collections and a variety of genetic markers useful in identifying individuals, researchers have delineated the spatial extent and temporal persistence of Amanita francheti, Laccaria bicolor, Suillus pungens, Pisolithus tinctorius, Hebeloma cylindrosporum, Lactarius xanthogalactus, Russula cremoricolor,

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1865

Fig. 4 Different views of community structure provided by species abundance vs. frequency. Relative abundance (black) and relative frequency (white) of mycorrhizae are shown for data from Pinus muricata/Pseudotsuga menziesii and Pinus muricata sites of Horton & Bruns (1998, a and b) and Gardes & Bruns (1996a, c and d), respectively. Species data are ranked by abundance as in the original papers. The top graphs show relative abundance and frequency separately, bottom shows relative abundance and relative frequency summed into an importance value. Relative abundance = tips or dry weight of a species/total tips or dry weight for all species. Absolute frequency = number of samples in which a species occurs/total number of samples. Relative frequency = absolute frequency of individual species/sum of absolute frequencies for all species. 1, Tomentella sublilacina; 2, Thelephoroid 2; 3, Russula amoenolens; 4, R. xerampelina-like 1; 5, Rhizopogon parksii; 6, Amanita gemmata/pantherina; 7, Cortinarioid-1; 8, Cenococcum geophilum; 9, Amanita muscaria; 10, unknown ascomycete; 11, Laccaria amethysteo-occidentalis; 12, Amanitoid; 13, Thelephoroid-3; 14, Xerocomus chrysenteron; 15, Thelephoroid-4; 16, Russuloid-2; 17, Boletoid-1; 18, Amanita francheti; 19, Clavulina sp.; 20, Cortinariod-2; 21, Rhizopogon subcaerulescens; 22, Boletoid-3; 23, Suillus pungens; 24, Russula xerampelina-like 2; 25, boletoid-2; 26, suilloid.

relatively small number of samples. Either a frequency or abundance view alone would not have revealed these patterns. Other species, such as Tomentella sublilacina (1), Russula amoenolens (3), and R. xerampelina-like 1 (4), were both frequent and abundant, and these species behaved this way at both sites. In these cases either an abundance or a frequency comparison would have identied these as dominant species, but presenting an importance value using relative dominance and relative frequency (Fig. 5b, 5d) provides an effective presentation for the contribution of each species in the community (Mueller-Dombois & Ellenberg 1974).

How much to sample


In stands with a high level of diversity, we may never be able to sample enough to detect treatment effects on the rarest species. Thus, with the current technology, it may be most appropriate to accept this and settle for analyses
2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

concerning only the dominant species; much of the data presented above highlights this approach. It will be benecial to reduce the number of root tips collected per sample to a manageable number, and then increase the number of samples taken in studies conducted in complex communities (Gehring et al. 1998; Jonsson et al. 1999a,b). It is also helpful to gain some concept of the level of diversity in a community prior to developing a sampling approach for a particular study site (Gehring et al. 1998). In ecosystems where EM fungal diversity is expected to be low or where species are expected to be more uniformly distributed, the problem is not as acute. This can be expected in relative harsh environments such as sites that have been recently disturbed (Baar et al. 1999), those that are undergoing primary succession (Gehring et al. 1998), or those that involve introductions of EM communities into non-EM habitats such as Pinus radiata plantations in paramo grasslands of Ecuadorian Andes (Chapela et al. 2001).

1866 T. R . H O RT O N and T. D . B R U N S
Over sort morphotypes Extract DNA PCR amplification

Tool

RFLP matching

Phylogenetic sequence analyses

Plant

Fungus Other genes such as large and small subunits of the nuclear and mitochondrial rRNA genes, protein genes Ribosomal database search, alignment and phylogenetic analysis Nuclear 5.8s rRNA gene

Region of Interest

Small portion of 28s rRNA

ITS

Portion of mitochondrial large rRNA gene

ITS

Analysis

Visual match

Visual match or cluster analysis

Alignment and phylogenetic analysis

Blast search, alignment, and phylogenetic analysis

Alignment and phylogenetic analysis

1 objective

Identify plant Group identical types family (genus)

Identify basidiomycete family group (genera)

Identify species group (species)

Identify genera & family-sized groups

Avoid analyzing non-target ITS sequences

2 objective

Identify species

Narrow search image for species ID

Narrow search image for species ID

Narrow search image for species ID

Identify phylum (fungus subdivision)

Fig. 5 Flowchart of a combined morphological-molecular typing approach to identify EM fungi from root tips. Rapid sorting of EM is conducted utilizing morphological characters, visible under a dissecting microscope: colour, presence/absence of extramatrical hyphae, branching pattern, presence/absence of setae or cystidia. EM from each soil sample are sorted into as many groups as possible at the expense of splitting species into multiple morphotypes. Each soil sample is analysed separately to further avoid lumping of species at the morphotyping step. The sorting of EM should be completed within two weeks after removal from the eld to maintain high quality DNA for PCR amplication. To identify EM types, two general techniques are employed, RFLP-matching and phylogenetic sequence analysis. For those fungi that remain unknown after comparison to a database of ITS RFLPs generated from sporocarp DNA, phylogenetic analysis of various gene regions provides increased taxonomic information from the phylum to the species group level. Identical types within and across soil samples are grouped after RFLP analysis and the data are then adjusted accordingly.

How much to morphotype


At one extreme, detailed descriptions of each morphotype can be generated whereby identical types can be recognized by morphology across samples (Shaw et al. 1993; Smith et al. 1995; Visser 1995; Helm et al. 1996; Luoma et al. 1997; Simard et al. 1997a,b; Goodman & Trofymow 1998a,b; Kranabetter & Wylie 1998; Qian et al. 1998; Massicotte et al. 1999). This can amount to a tremendous effort; Luoma et al. (1997) characterized 198 morphological types from 200 soil samples. Unfortunately, despite the effort, the species identication for all but a few types is left unclear when morphology is used exclusively. At the other extreme, one or a few root tips for ITSRFLP identication from each soil sample can be selected without the morphotyping step (Gehring et al. 1998; Jonsson et al. 1999c). This can be very fast and is especially useful in settings were multiple species are known to form mycorrhizae with a similar morphology (e.g. smooth brown, Jonsson et al. 1999c). It is probably the method of choice for questions involving large spatial scales, but resolution will be essentially limited to frequency changes in the dominant species. A combined approach using rapid sorting by EM morphology followed by molecular identication can be employed (Fig. 5; Erland 1995; Mehmann et al. 1995; Gardes & Bruns 1996b; Dahlberg et al. 1997; Krn & Nylund 1997; Kernaghan et al. 1997; Pritsch et al. 1997; Jonsson et al.

1999a,c). This can be coupled with a complete sampling of roots within a soil sample, which has the advantage of increasing the likelihood of nding rare types (Gardes & Bruns 1996a; Horton & Bruns 1998; Horton et al. 1999; Taylor & Bruns 1999; Bidartondo et al. 2000). However, if one uses a crude morphotyping approach, it should be limited to sorting differences within soil samples rather than using it to relate types between samples. Identical types encountered in separate cores, or that were inadvertently separated within a core, are numerically combined only after RFLP analysis. The two reasons for structuring the morphotype analysis in this way are that direct visual comparisons are convenient for all morphotypes within a soil sample, and the spatial scale will limit diversity when crude morphological differences are used as the primary distinction. This lessens the lumping of similar looking types, such as the brown type in Erland (1995) or the white types in Jonsson et al. (2000), and it results in multiple RFLP replicates of common types. Furthermore, if morphotypes are later found to contain two or more RFLP types, it is relatively easy to analyse additional archived samples of the morphotype and correct the error by estimating the proportion of each RFLP type found with a single morphotype (Horton & Bruns 1998; Jonsson et al. 1999a,b). Using a combination of rapid morphological sorting, RFLP matching, and sequence analysis has proven a successful approach for identifying EM fungi from root tips.
2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1867 On average, our laboratory groups have identied EM types to the species level for about 40% of our EM by biomass, with another 40% that were sorted to species group level by RFLP and then identied to family group through sequence analysis (Table 1). Similar levels of identication were also achieved by Peter et al. (2001) using the same approach. cation outlined here with research directed towards understanding the functional roles of various species is required. To date, most of the physiological and autecological studies have been focused on a handful of fungi that grow well in culture or that are readily collected as sporocarps. Now that we have identied a different set of dominant taxa, increased effort is needed to understand their unique roles and behaviour.

Mycelial view back to quantication by molecules?


The active soil mycelium has clear functional signicance, as demonstrated by the acquisition and transfer of nutrients throughout the thallus and to associated plants (Melin & Nilsson 1950, 1953; Finlay & Read 1986; Finlay et al. 1989; Bending & Read 1995a,b; Perez-Moreno & Read 2000). Yet, the distribution and abundance of fungal mycelium in the soil remains largely undocumented in the eld. It seems likely that the mycelial view of the EM community will be different from both root and fruitbody views. While we know that some fungi form hyphal cords, rhizomorphs, fans, or mats, and others do not, data are lacking for most species, and the relationship between mycelial investment and root dominance is unknown. If quantifying species by mycelia is the goal, then the techniques used will likely become more similar to those used in bacterial systems. Obvious approaches would include: (i) denaturing gradient gel electrophoresis, temperature gradient gel electrophoresis, or terminally labelled characterization of amplied portions of the rDNA (Muyzer & Smalla 1998; Moeseneder et al. 1999); and (ii) quantication of rRNA via hybridization probes including microarrays and in situ hybridization (Bruns & Gardes 1993; Amann 1995; Zheng et al. 1996) coupled with microscopic quantication of mycelium. Because the rst set of techniques involve PCR amplication, there will be inherent biases that make quantication approximate at best; nevertheless these kinds of approaches have been very useful in bacterial systems for identifying the dominant species in a system. The second two approaches, typically target the unamplied ribosomes, and so would not have PCR biases, but they would require of set of well characterized oligonucleotide probes that have not been developed yet for most EM fungi.

Acknowledgements
We thank the following for providing helpful comments on earlier drafts of this paper: Martin Bidartondo, Lee Taylor, and Annette Kretzer. We thank Erik Lilleskov for critical advice and help with the species/area curves. Funding was provided in part by NSF grant DEB9815262 to T. D. Bruns and NRI Competitive Grants Program/USDA award 99351077843 and support from the US Forest Service, PNW Research Station to T. R. Horton.

References
Aanen DK, Kuyper TW, Boekhout T, Hoekstra RF (2000) Phylogenetic relationships in the genus Hebeloma based on ITS1 and 2 sequences, with special emphasis on the Hebeloma crustiliniforme complex. Mycologia, 92, 269281. Agerer R (198796) Colour Atlas of Ectomycorrhizae. Einhorn-Verlag Eduard Dietenberger, Schwbisch Gmnd. Allen MF (1991) The Ecology of Mycorrhizae Cambridge University Press, Cambridge. Amann RI (1995) Fluorescently labelled, rRNA-targeted oligonucleotide probes in the study of microbial ecology. Molecular Ecology, 4, 543554. Anderson IC, Chambers SM, Cairney JWG (1998) Use of molecular methods to estimate the size and distribution of mycelial individuals of the ectomycorrhizal basidiomycete Pisolithus tinctorius. Mycological Research, 102, 295 300. Arnolds E (1988) The changing macromycete ora in the Netherlands. Transactions of the British Mycological Society, 90, 391406. Baar J, Horton TR, Kretzer AM, Bruns TD (1999) Mycorrhizal colonization of Pinus muricata from resistant propagules after a stand replacing wildre. New Phytologist, 143, 409 418. Bending GD, Read DJ (1995a) The structure and function of the vegetative mycelium of ectomycorrhizal plants V. Foraging behaviour and translocation of nutrients from exploited litter. New Phytologist, 130, 401409. Bending GD, Read DJ (1995b) The structure and function of the vegetative mycelium of ectomycorrhizal plants VI. Activities of nutrient mobilising enzymes in birch litter colonised by Paxillus involutus (Fr.) Fr. New Phytologist, 130, 411 417. Bidartondo M, Kretzer AM, Bruns TD (2000) High root concentration and uneven ectomycorrhizal diversity near Sarcodes sanguinea (Ericaceae): a cheater that stimulates its victims? American Journal of Botany, 87, 17831788. Bjrkman E (1960) Monotropa hypopitys L. an epiparisite on tree roots. Physiologia Plantarum, 13, 308327. Bledsoe CS, Tennyson K, Lopushinsky W (1982) Survival and growth of outplanted Douglas-r seedlings inoculated with mycorrhizal fungi. Canadian Journal of Forest Research, 12, 720723.

Concluding remarks
We have come a long way from the black-box approach in terms of identifying the basic structure of EM communities. However, methodological improvements, particularly in the use of sequence-level characterization, need to be employed to move beyond the large number of unknown RFLP-types. We also need ways to increase sample sizes so that broad-scale questions can be addressed more efciently. Linking the detailed efforts in identi 2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

1868 T. R . H O RT O N and T. D . B R U N S
Bonello P, Bruns TD, Gardes M (1998) Genetic structure of a natural population of the ectomycorrhizal fungus Suillus pungens. New Phytologist, 138, 533 542. Brandrud TE (1995) The effects of experimental nitrogen addition on the ectomycorrhizal fungus ora in an oligotrophic spruce forest at Grdshjn, Sweden. Forest Ecology and Management, 71, 111122. Bruns TD, Gardes M (1993) Molecular tools for the identication of ectomycorrhizal fungi: taxon-specic oligonucleotide probes for suilloid fungi. Molecular Ecology, 2, 233242. Bruns TD, Szaro TM, Gardes M et al. (1998) A sequence database for the identication of ectomycorrhizal basidiomycetes by phylogenetic analysis. Molecular Ecology, 7, 257272. Bruns TD, Vilgalys R, Barns SM et al. (1992) Evolutionary relationships within the fungi: Analysis of nuclear small subunit rRNA sequences. Molecular Phylogenetics and Evolution, 3, 231241. Burnham KP, Overton WS (1979) Robust estimation of population size when capture probabilities vary among animals. Ecology, 60, 927 936. Byrd KB, Parker VT, Vogler DR, Cullings KW (2000) The inuence of clear-cutting on ectomycorrhizal fungus diversity in a lodgepole pine (Pinus contorta) stand, Yellowstone National Park, Wyoming, and Gallatin National Forest, Montana. Canadian Journal of Botany, 78, 149 156. Camacho FJ, Gernandt DS, Liston A et al. (1997) Endophytic fungal DNA, the source of contamination in spruce needle DNA. Molecular Ecology, 6, 983 987. Chambers SM, Sharples JM, Cairney JWG (1998) Towards a molecular identication of the Pisonia mycobiont. Mycorrhiza, 7, 319 321. Chapela IH, Osher LJ, Horton TR, Henn MR (2001) Ectomycorrhizal fungi introduced with exotic pine plantations induce soil carbon depletion. Soil Biology and Biochemistry, in press. Chapin FS III, Shaver GR (1985) Individualistic growth response of tundra plant species to environmental manipulations in the eld. Ecology, 66, 564 576. Cullings KW (1992) Design and testing of a plant-specic PCR primer for ecological and evolutionary studies. Molecular Ecology, 1, 233 240. Cullings KW, Szaro TM, Bruns TD (1996) Evolution of extreme specialization within a lineage of ectomycorrhizal epiparasites. Nature, 379, 63 66. Cullings KW, Vogler DR (1998) A 5.8S nuclear ribosomal RNA gene sequence database: applications to ecology and evolution. Molecular Ecology, 7, 919 923. Cullings KW, Vogler DR, Parker VT, Finley SK (2000) ectomycorrhizal specicity patterns in a mixed Pinus contorta/Picea engelmannii forest in Yellowstone National Park. Applied and Environmental Microbiology, 66, 49884991. Dahlberg A, Jonsson L, Nylund J-E (1997) Species diversity and distribution of biomass above and below ground among ectomycorrhizal fungi in an old-growth Norway spruce forest in south Sweden. Canadian Journal of Botany, 75, 13231335. Danielson RM (1984) Ectomycorrhizal associations in jack pine stands in northeastern Alberta. Canadian Journal of Botany, 62, 932 939. Danielson RM, Visser S (1989) Host response to inoculation and behavior of introduced and indigenous ectomycorrhizal fungi of jack pine grown on oil-sands tailings. Canadian Journal of Forest Research, 19, 1412 1421. De La Bastide PY, Kropp BR, Piche Y (1994) Spatial distribution and temporal persistence of discrete genotypes of the ectomycorrhizal fungus Laccaria bicolor (Maire) Orton. New Phytologist, 127, 547556. Eberhardt U, Oberwinkler F, Verbeken A et al. (2000) Lactarius ectomycorrhizae on Abies alba: morphological description, molecular characterization, and taxonomic remarks. Mycologia, 92, 860873. Egger KN (1995) Molecular analysis of ectomycorrhizal fungal communities. Canadian Journal of Botany, 73, S1415 S1422. Erland S (1995) Abundance of Tylospora brillosa ectomycorrhizas in a South Swedish spruce forest measured by RFLP analysis of the PCR-amplied rDNA ITS region. Mycological Research, 99, 14251428. Farmer DJ, Sylvia DM (1998) Variation in the ribosomal DNA internal transcribed spacer of a diverse collection of ectomycorrhizal fungi. Mycological Research, 102, 859 865. Finlay RD, Ek H, Odham G, Sderstrom B (1989) Uptake, translocation and assimilation of nitrogen from 15N-labelled ammonium and nitrate sources by intact ectomycorrhizal systems of Fagus sylvatica infected with Paxillus involutus. New Phytologist, 120, 4755. Finlay RD, Read DJ (1986) The structure and function of the vegetative mycelium of ectomycorrhizal plants II. The uptake and distribution of phosphorus by mycelial strands interconnecting host plants. New Phytologist, 103, 157165. Flynn D, Newton AC, Ingleby K (1998) Ectomycorrhizal colonisation of Sitka spruce (Picea sitchensis (Bong.) Carr) seedlings in a Scottish plantation forest. Mycorrhiza, 7, 313 317. Gardes M, Bruns TD (1993) ITS primers with enhanced specicity for basidiomycetes application to the identication of mycorrhizae and rusts. Molecular Ecology, 2, 113 118. Gardes M, Bruns TD (1996a) Community structure of ectomycorrhizal fungi in a Pinus muricata forest: above- and below-ground views. Canadian Journal of Botany, 74, 1572 1583. Gardes M, Bruns TD (1996b) ITS-RFLP matching for identication of fungi. Methods in Molecular Biology, 50, 177186. Gardes M, White TJ, Fortin JA, Bruns TD, Taylor JW (1991) Identication of indigenous and introduced symbiotic in ectomycorrhizae by amplication of the nuclear and mitochondrial ribosomal DNA. Canadian Journal of Botany, 69, 180 190. Gargas A, Depriest PT, Grube M, Tehler A (1995) Multiple origins of lichen symbioses in fungi suggested by SSU rDNA phylogeny. Science, 268, 14921495. Gehring CA, Theimer TC, Whitham TG, Keim P (1998) Ectomycorrhizal fungal community structure of pinyon pines growing in two environmental extremes. Ecology, 79, 1562 1572. Goodman DM, Durall DM, Trofymow JA, Berch SM, eds (1996 98) A Manual of Concise Descriptions of North American Ectomycorrhizae. Mycologue Publications and Canada-B.C. Forest Resource Development Agreement, Canadian Forest Service, Victoria, B.C. Goodman DM, Trofymow JA (1998a) Comparison of communities of ectomycorrhizal fungi in old-growth and mature stands of Douglas-r at two sites on southern Vancouver island. Canadian Journal of Forest Research, 28, 574581. Goodman DM, Trofymow JA (1998b) Distribution of ectomycorrhizas in microhabitats in mature and old-growth stands of Douglas-r on southeastern Vancouver island. Soil Biology and Biochemistry, 30, 21272138. Gyrta H, Debaud J-C, Effosse A, Gay G, Marmeisse R (1997) Finescale structure of populations of the ectomycorrhizal fungus Hebeloma cylindrosporum in coastal sand dune forest ecosystems. Molecular Ecology, 6, 353364. 2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1869
Hagerman SM, Jones MD, Bradeld GE, Gillespie M, Durall DM (1999) Effects of clear-cut logging on the diversity and persistence of ectomycorrhizae at a subalpine forest. Canadian Journal of Forest Research, 29, 124 134. van der Heijden MGA, Klironomos JN, Ursic M et al. (1998) Mycorrhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature, 396, 6972. Helm DJ, Allen EB, Trappe JM (1996) Mycorrhizal chronosequence near Exit Glacier, Alaska. Canadian Journal of Botany, 74, 1496 1506. Henrion B, Le Tacon F, Martin F (1992) Rapid identication of genetic variation of ectomycorrhizal fungi by amplication of ribosomal RNA genes. New Phytologist, 122, 289298. Hibbett DS, Gilbert LB, Donohue MJ (2000) Evolutionary instability of ectomycorrhizal symbioses in basidiomycetes Nature, 407, 506 508. Hibbett DS, Pine EM, Langer E et al. (1997) Evolution of gilled mushrooms and puffballs inferred from ribosomal DNA sequences. Proceedings of the National Academy of Sciences of the USA, 94, 12002 12006. Horton TR, Bruns TD (1998) Multiple host fungi are the most frequent and abundant ectomycorrhizal types in a mixed stand of Douglas r (Pseudotsuga menziesii (Mirb.) Franco) and bishop pine (Pinus muricata D. Don). New Phytologist, 139, 331339. Horton TR, Bruns TD, Parker VT (1999) Ectomycorrhizal fungi associated with Arctostaphylos contribute to Pseudotsuga menziesii establishment. Canadian Journal of Botany, 77, 93102. Ingleby K, Mason PA, Last FT, Fleming LV (1990) Identication of Ectomycorrhizas. HMSO, London. Jacobson KM, Miller OK, Turner BJ (1993) Randomly amplied polymorphic DNA markers are superior to somatic incompatibility tests for discriminating genotypes in natural populations of the ectomycorrhizal fungus Suillus granulatus. Proceedings of the National Academy of Sciences of the USA, 90, 91599163. Jonsson L, Dahlberg A, Brandrud T-E (2000) Spatiotemporal distribution of an ectomycorrhizal community in an oligotrophic Swedish Picea abies forest subjected to experimental nitrogen addition: above- and below-ground views. Forest Ecology and Management, 132, 143 156. Jonsson L, Dahlberg A, Nilsson M-C, Krn O, Zackrisson O (1999a) Continuity of ectomycorrhizal fungi in self-regenerating boreal Pinus sylvestris forests studied by comparing mycobiont diversity on seedlings and mature trees. New Phytologist, 142, 151162. Jonsson L, Dahlberg A, Nilsson M-C, Zackrisson O, Krn O (1999b) Ectomycorrhizal fungal communities in late-successional Swedish boreal forests, and their composition following wildre. Molecular Ecology, 8, 205 215. Jonsson T, Kokalj S, Finlay R, Erland S (1999c) Ectomycorrhizal community structure in a limed spruce forest. Mycological Research, 103, 501 508. Krn O, Hogberg N, Dahlberg A et al. (1997) Inter- and intraspecic variation in the ITS region of rDNA of ectomycorrhizal fungi in Fennoscandia as detected by endonuclease analysis. New Phytologist, 136, 313 325. Krn O, Nylund J-E (1997) Effects of ammonium sulphate on the community structure and biomass of ectomycorrhizal fungi in a Norway spruce stand in southwestern Sweden. Canadian Journal of Botany, 75, 1628 1642. Kernaghan G, Currah RS, Bayer RJ (1997) Russulaceous ectomycorrhizae of Abies lasiocarpa and Picea engelmannii. Canadian Journal of Botany, 75, 1843 1850. 2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871 Kranabetter JM, Wylie T (1998) Ectomycorrhizal community structure across forest openings on naturally regenerated western hemlock seedlings. Canadian Journal of Botany, 76, 189 196. Kretzer AM, Bidartondo MI, Grubisha LC et al. (2000a) Regional specialization of Sarcodes sanguinea (Ericaceae) on a single fungal symbiont from the Rhizopogon ellenae (Rhizopogonaceae) species complex. American Journal of Botany, 87, 1778 1782. Kretzer AM, Li Y, Szaro T, Bruns TD (1996) Internal transcribed spacer sequences from 38 recognized species of Suillus sensu lato: phylogenetic and taxonomic implications. Mycologia, 88, 776 785. Kretzer AM, Molina R, Spatafora JW (2000b) Microsatellite markers for the ectomycorrhizal basidiomycete Rhizopogon vinicolor. Molecular Ecology, 9, 11901191. Kropp BR, Trappe JM (1982) Ectomycorrhizal fungi of Tsuga heterophylla. Mycologia, 74, 479488. Landvik S, Egger KN, Schumacher T (1997) Towards a subordinal classication of the Pezizales (Ascomycota): Phylogenetic analyses of SSU rDNA sequences. Nordic Journal of Botany, 17, 403418. Lanfranco L, Perotto S, Bonfante P (1998) Applications of PCR for studying the biodiversity of mycorrhizal fungi. In: Applications of PCR in Mycology (eds Bridge PD, Arora DK, Reddy CA, Elander RP), pp. 107124. CAB International, UK. Larsen M (1968) Tomentelloid fungi of North America. State University College of Forestry at Syracuse University. Technical Publication, 93, p. 157. LePage BA, Currah RS, Stockey RA, Rothwell GW (1997) Fossil ectomycorrhizae from the middle Eocene. American Journal of Botany, 84, 410412. Lilleskov EA, Fahey TJ, Horton TR, Lovett GM (2002) Nitrogen deposition and ectomycorrhizal fungal communities: a belowground view from Alaska. Ecology, 83, in press. Liu YJ, Rogers SO, Ammirati JF (1997) Phylogenetic relationships in Dermocybe and related Cortinarius taxa based on nuclear ribosomal DNA internal transcribed spacers. Canadian Journal of Botany, 75, 519532. Liu YJ, Rogers SO, Ammirati JF, Keller G (1995) Dermocybe, Section Sanguineae: a look at species within the sanguinea complex. Sydowia, 10, 142154. Luoma DL (1991) Annual changes in seasonal production of hypogeous sporocarps in Oregon Douglas-r forests. In: Wildlife and Vegetation of Unmanaged Doulgas-Fir Forests (eds Ruggieros LF, Aubry KB, Carey AB, Huff MH), pp. 83 89. USDA Forest Service General Technical Report, Portland. Luoma DL, Eberhart JL, Amaranthus MP (1997) Biodiversity of ectomycorrhizal types from southwest Oregon. In: Conservation and Management of Native Plants and Fungi (eds Kaye TN, Liston A, Love RM, Luoma DL, Meinke RJ, Wilson MV), pp. 249 253. Corvallis, Native Plant Society of Oregon, Oregon. Mahmood S, Finlay R, Erland S (1999) Effects of repeated harvesting of forest residues on the ectomycorrhizal community in a Swedish spruce forest. New Phytologist, 142, 557 585. Massicotte HB, Molina R, Tackaberry LE, Smith JE, Amaranthus MP (1999) Diversity and host specicity of ectomycorrhizal fungi retrieved from three adjacent forest sites by ve host species. Canadian Journal of Botany, 77, 1053 1076. McCune B, Mefford MJ (1999) PC-ORD. Multivariate Analysis of Ecological Data, Version 4. MjM Software Design, Gleneden Beach, Oregon, USA. McKendrick SL, Leake JR, Taylor DL, Read DJ (2000) Symbiotic germination and development of myco-hetertrophic plants in nature: transfer of carbon from ectomycorrhizal Salix repens and

1870 T. R . H O RT O N and T. D . B R U N S
Betula pendula to the orchid Corallorhiza trida through shared hyphal connections. New Phytologist, 145, 539548. Mehmann B, Egli S, Bravs GH, Brunner I (1995) Coincidence between molecularly or morphologically classied ectomycorrhizal morphotypes and fruitbodies in a Spruce forest. In: Biotechnology of Ectomycorrhizae: Molecular Approaches (eds Stocchi VP, Bonfante P, Nuti M), pp. 4152. Plenum Press, London. Melin E, Nilsson H (1950) Transfer of radioactive phosphorus to pine seedlings by means of mycorrhizal hyphae. Physiologia Plantarum, 3, 88 92. Melin E, Nilsson H (1953) Transfer of labeled nitrogen from glutamic acid to pine seedlings through the mycelium of Boletus variegatus (S.W.) Fr. Nature, 171, 134. Methvyn A, Hughes KW, Petersen RH (2000) Flammulina RFLP patterns identify species and show biogeographical patterns within species. Mycologia, 92, 1064 1070. Moeseneder MM, Arrieta JM, Muyzer G et al. (1999) Optimization of terminal-restriction fragment length polymorphism analysis for complex marine bacterioplankton communities and comparison with denaturing gradient gel electrophoresis. Applied and Environmental Microbiology, 65, 35183525. Molina R, Massicotte H, Trappe JM (1992) Specicity phenomena in mycorrhizal symbiosis: community-ecological consequences and practical implications. In: Mycorrhizal Functioning: an Integrated Plant-Fungal Process (ed. Allen MF), pp. 357423. Chapman & Hall, London. Molina R, Trappe JM (1982) Lack of mycorrhizal specicity by the ericaceous hosts Arbutus menziesii and Arctostaphylos uva-ursi. New Phytologist, 90, 495 509. Moncalvo JM, Lutzoni FM, Rehner SA et al. (2000) Phylogenetic relationships of agaric fungi based on nuclear large subunit ribosomal DNA sequences Systematic Biology, 49, 278305. Mueller-Dombois D, Ellenberg H (1974) Aims and Methods of Vegetation Ecology. John Wiley and Sons, New York. Mullis KB, Faloona FA (1987) Specic synthesis of DNA in vitro via a polymerase-catalyzed chain reaction. Methods in Enzymology, 155, 335 350. Muyzer G, Smalla K (1998) Application of denaturing gradient gel electrophoresis (DGGE) and temperature gradient gel electrophoresis (TGGE) in microbial ecology. Antonie Van Leeuwenhoek, 73, 127141. Newman EI (1988) Mycorrhizal links between plants: their functioning and ecological signicance. Advances in Ecological Research, 18, 243 270. Norman JE, Egger KN (1996) Phylogeny of the genus Plicaria and its relationship to Peziza inferred from ribosomal DNA sequence analysis. Mycologia, 88, 986995. Norman JE, Egger KN (1999) Molecular phylogenetic analysis of Peziza and related genera. Mycologia, 91, 820829. Nylund JE, Dahlberg A, Hogberg N etal. (1995) Methods for studying species composition of mycorrhizal fungal communities in ecological studies and environmental monitoring. In: Biotechnology of Ectomycorrhizae: Molecular Approaches (eds Stocchi V, Bonfante P, Nuti M), pp. 229 239. Plenum Press, New York. ODonnell K, Cigelnik E, Weber NS, Trappe JM (1997) Phylogenetic relationships among ascomycetous trufes and the true and false morels inferred from 18S and 28S ribosomal DNA sequence analysis. Mycologia, 89, 4865. Olsen GJ, Overbeek R, Larsen N et al. (1993) The ribosomal database project. Nucleic Acids Research, 20, 21992200. Palmer MW (1991) Estimating species richness: the second-order jackknife reconsidered. Ecology, 72, 15121513. Perez-Moreno J, Read DJ (2000) Mobilization and transfer of nutrients from litter to tree seedlings via the vegetative mycelium of ectomycorrhizal plants. New Phytologist, 145, 301 309. Perry DA, Amaranthus MP, Borchers JG, Borchers SL, Brainerd RE (1989) Bootstrapping in ecosystems. Bioscience, 39, 230 237. Peter M, Ayer F, Egli S (2001) Nitrogen addition in Norway spruce stand altered macromycete sporocarp production and belowground ectomycorrhizal species composition measured by PCRRFLP analysis of the ribosomal ITS-region. New Phytologist, 149, 311324. Pine E, Hibbett DS, Donoghue MJ (1999) Phylogenetic relationships of cantharelloid and clavarioid homobasidiomycetes based on mitochondrial and nuclear rDNA sequences. Mycologia, 91, 944963. Pirozynski KA, Dalp Y (1989) Geological history of Glomaceae with particular reference to mycorrhizal symbiosis. Symbiosis, 7, 136. Pritsch K, Boyle H, Munch JC, Buscot F (1997) Characterization and identication of black alder ectomycorrhizas by PCR-RFLP analyses of the rDNA internal transcribed spacer (ITS). New Phytologist, 137, 357369. Pritsch K, Munch J-C, Buscot F (2000) Identication and differentiation of mycorrhizal isolates of black alder by sequence analysis of the ITS region. Mycorrhiza, 10, 87 93. Qian XM, Kottke I, Oberwinkler F (1998) Inuence of liming and acidication on the activity of the mycorrhizal communities in a Picea abies (L.) Karst. stand. Plant and Soil, 199, 99 109. Read DJ (1991) Mycorrhizas in ecosystems. Experentia, 47, 376 391. Redecker D, Hijri M, Dulieu H, Sanders IR (1999) Phylogenetic analysis of a dataset of fungal 5.8S rDNA sequences shows that highly divergent copies of internal transcribed spacers reported from Scutellospora castanea are of ascomycete origin. Fungal Genetics and Biology, 28, 238244. Redecker D, Szaro TM, Bowman RJ, Bruns TD (2001) Small genets of Lactarius xanthogalactus, Russula cremoricolor and Amanita francheti in late stage ectomycorrhizal successions. Molecular Ecology, 10, 10251034. Sawyer NA, Chambers SM, Cairney JWG (1999) Molecular investigation of genet distribution and genetic variation of Cortinarius rotundisporus in eastern Australian sclerophyll forests. New Phytologist, 142, 561568. Selosse M-A, Jacquot D, Bouchard D, Martin F, Le Tacon F (1998a) Temporal persistence and spatial distribution of an American inoculant strain of the ectomycorrhizal basidiomycete Laccaria bicolor in a French forest plantation. Molecular Ecology, 7, 561 573. Selosse M-A, Martin F, Bouchard D, Le Tacon F (1999) Structure and dynamics of experimentally introduced and naturally occurring Laccaria sp. Discrete genotypes in a Douglas r plantation. Applied and Environmental Microbiology, 65, 2006 2014. Selosse M-A, Martin F, Le Tacon F (1998b) Survival of an introduced ectomycorrhizal Laccaria bicolor strain in a European forest plantation monitored by mitochondrial by mitochondrial ribosomal DNA analysis. New Phytologist, 140, 753 761. Shaw PJA, Dighton J, Poskitt J (1993) Studies on the mycorrhizal community infecting trees in the Liphook Forest fumigation experiment. Agriculture Ecosystems and Environment, 47, 185 191. Simard SW, Molina R, Smith JE, Perry DA, Jones MD (1997a) Shared compatibility of ectomycorrhizae on Pseudotsuga menziesii and Betula papyrifera seedlings grown in mixture in soils from southern British Columbia. Canadian Journal of Forest Research, 27, 331342. 2001 Blackwell Science Ltd, Molecular Ecology, 10, 1855 1871

M O L E C U L A R A P P R O A C H E S F O R S T U D I E S O F E M E C O L O G Y 1871
Simard SW, Perry DA, Jones MD et al. (1997b) Net transfer of carbon between ectomycorrhizal tree species in the eld. Nature, 388, 579 582. Simon L, Bousquet J, Levesque RC, LaLonde M (1993) Origin and diversication of endomycorrhizal fungi and coincidence with vascular land plants. Nature, 363, 6769. Smith JE, Molina R, Perry DA (1995) Occurrence of ectomycorrhizas on ericeous and coniferous seedlings grown in soils from the Oregon Coast Range. New Phytologist, 129, 7381. Smith SE, Read DJ (1997) Mycorrhizal Symbioses 2nd edn. Academic Press, London. Stendell ER, Horton TR, Bruns TD (1999) Early effects of prescribed re on the structure of the ectomycorrhizal fungus community in a Sierra Nevada ponderosa pine forest. Mycological Research, 103, 1353 1359. Sylvia DM, Jarstfer AG (1997) Distribution of mycorrhiza on competing pines and weeds in a southern pine plantation. Soil Science Society of America Journal, 61, 139144. Taylor AFS, Alexander IJ (1989) Demography and population dynamics of ectomycorrhizas of Sitka spruce fertilised with N. Agriculture Ecosystems and Environment, 28, 493496. Taylor DL, Bruns TD (1997) Independent, specialized invasions of ectomycorrhizal mutualism by two nonphotosynthetic orchids. Proceedings of the National Academy of Sciences of the USA, 94, 4510 4515. Taylor DL, Bruns TB (1999) Community structure of ectomycorrhizal fungi in a Pinus muricata forest: minimal overlap between the mature forest and resistant propagule communities. Molecular Ecology, 8, 18371850. Tilman D (1994) Competition and biodiversity in spatially structured habitats. Ecology, 75, 2 16. Visser S (1995) Ectomycorrhizal fungal succession in jack pine stands following wildre. New Phytologist, 129, 389 401. Vitousek PM, Aber JD, Howarth RH et al. (1997) Human alteration of the global nitrogen cycle: Source and consequences. Ecological Applications, 7, 737750. Vrlstad T, Fossheim T, Schumacher T (2000) Picearhiza bicolorata the ectomycorrhizal expression of the Hymenoscyphus ericaea aggregate? New Phytologist, 145, 549563. White TJ, Bruns TD, Lee SB, Taylor JW (1990) Amplication and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a Guide to Methods and Applications (eds Innis MA, Gelfand DH, Sninsky JJ, White TJ), pp. 315 322. Academic Press, London. Zheng D, Alm EW, Stahl DA, Raskin L (1996) Characterization of universal small-subunit rRNA hybridization probes for quantitative molecular microbial ecology studies. Applied and Environmental Microbiology, 62, 45044513. Zhou Z, Makoto M, Hogetsu T (1999) Analysis of genetic structure of a Suillus grevillei population in a Larix kaempferi stand by polymorphism of inter-simple sequence repeat (ISSR). New Phytologist, 144, 5563.

The authors are generally interested in ectomycorrhizal fungi and their roles in plant community dynamics. Primary topics of interest include elucidating: (i) the population and community structure of EM fungi; (ii) the role of EM fungi in plant establishment following disturbances and during primary and secondary succession; and (iii) delineating the phylogenetic relationships of EM fungi.

2001 Blackwell Science Ltd, Molecular Ecology, 10, 18551871

Vous aimerez peut-être aussi